Mechanical Stretching-Induced Electron-Transfer Reactions and

Sep 26, 2017 - The ratio of the high to low conductance peak areas (including conductance traces with and without switching events (Supporting Figure ...
0 downloads 10 Views 2MB Size
Article Cite This: J. Am. Chem. Soc. 2017, 139, 14699-14706

pubs.acs.org/JACS

Mechanical Stretching-Induced Electron-Transfer Reactions and Conductance Switching in Single Molecules Yueqi Li,†,‡ Naomi L. Haworth,§ Limin Xiang,†,‡ Simone Ciampi,*,∥ Michelle L. Coote,*,§ and Nongjian Tao*,†,⊥ †

Center for Bioelectronics and Biosensors, Biodesign Institute, ‡School of Molecular Sciences, and ⊥School of Electrical, Energy and Computer Engineering, Arizona State University, Tempe, Arizona 85287-5801, United States § ARC Centre of Excellence for Electromaterials Science, Research School of Chemistry, Australian National University, Canberra, Australian Capital Territory 2601, Australia ∥ Department of Chemistry, Curtin University, Bentley, Western Australia 6102, Australia S Supporting Information *

ABSTRACT: A central idea in electron-transfer theories is the coupling of the electronic state of a molecule to its structure. Here we show experimentally that fine changes to molecular structures by mechanically stretching a single metal complex molecule via changing the metal−ligand bond length can shift its electronic energy levels and predictably guide electron-transfer reactions, leading to the changes in redox state. We monitor the redox state of the molecule by tracking its characteristic conductance, determine the shift in the redox potential due to mechanical stretching of the metal−ligand bond, and perform model calculations to provide insights into the observations. The work reveals that a mechanical force can shift the redox potential of a molecule, change its redox state, and thus allow the manipulation of single molecule conductance.



INTRODUCTION Electron-transfer reactions are responsible for many chemical and biological processes, including photovoltaic energy conversion, photosynthesis, and metabolic activities in living systems.1 Central to electron-transfer reaction theories2 is the idea that the electronic state of a molecule is coupled to the nuclear degrees of freedom of the molecule and its surrounding solvent, and thermal fluctuations create a favorable nuclear geometry that allows electron transfer to take place.3 Here we show via single molecule conductance measurements that mechanically stretching a redox molecule can help create the favorable geometry, shift the redox potential, and induce electron-transfer reactions. We measure electron transfer in a single metal complex molecule that is covalently connected to two electrodes4 while stretching the metal−ligand bond length. This allows us to detect mechanical stretching-induced switching of the metal ion between reduced and oxidized states and to establish experimentally a link between changes in nuclear geometry and redox potential. Our theoretical modeling of the system confirms that the mechanical distortion in the metal−ligand bonds is responsible for the observed shift in the redox potential. The work points to a way to examine the interplay between electronic and structural properties of molecules in electron-transfer reactions and to switch the conductance of a single molecule via mechanically changing its © 2017 American Chemical Society

redox state without bond rupture. This could lead to a mechanically controllable molecular switch.5



RESULTS AND DISCUSSIONS We study the electron-transfer reactions of a ferrocene compound by monitoring its conductance while mechanically stretching the molecule under electrochemical control in 0.1 M NaClO4 aqueous solution. Ferrocene can reversibly donate an electron to an electrode (become oxidized) and accept an electron from the electrode (become re-reduced) without bond rupture.6 The ferrocene compound studied here is 1,1′ferrocenyl diester (Fc-Lip, see Figure 1a for structure, Supporting Figures 1-2 for NMR characterization, and Supporting Methods for synthesis protocols), which consists of a ferrocene redox center connected to two linker groups that can bind to the gold electrodes. As a control experiment, we also study 1,4-phenylene bis(5-(1,2-dithiolan-3-yl) pentanoate) (Lip-ctr, see Supporting Methods for synthesis protocols) which is similar to Fc-Lip but lacks the ferrocene redox center (Figure 1a and Supporting Figures 3−4). We measure the redox states of Fc-Lip from its characteristic conductance with the scanning tunneling microscope (STM) break junction Received: August 3, 2017 Published: September 26, 2017 14699

DOI: 10.1021/jacs.7b08239 J. Am. Chem. Soc. 2017, 139, 14699−14706

Article

Journal of the American Chemical Society

Figure 1. (a) Structures of Fc-Lip and Lip-ctr. (b) Experimental setup showing a redox molecule bridged between a gold STM tip and gold substrate for interrogating the redox state of the molecule via conductance measurement and for stretching the molecule. The tip and substrate potentials are controlled with respect to a reference electrode (silver wire as quasi-reference, potential calibrated in Supporting Figure 11) in 0.1 M NaClO4 solution. A small bias (30 mV) is maintained between the tip and substrate electrodes for conductance measurement. A platinum wire (not shown) is used as a counter electrode for stable potential control. (c) Cyclic voltammogram of Fc-Lip on a gold surface showing well-defined oxidation and reduction peaks at ∼0.4 V vs Ag/AgCl, characteristic of fully reversible redox reaction (potential sweeping rate: 100 mV/s). (d) Conductance of FcLip vs overpotential (E − Eox/red, where E is the potential and Eox/red is the redox potential), revealing a low conductance reduced state, a high conductance oxidized states, and switching between the two states (potential sweeping rate: 1 V/s).

technique7,8 (details in Materials and Methods section), where the gold tip and gold substrate serve as the two electrodes to change the geometry of the molecule via mechanical stretching and also to determine its conductance. During the measurement, we control the potential of the molecules with respect to a reference electrode (Figure 1b). The redox reaction of Fc-Lip immobilized on the electrodes (see Supporting Methods for immobilization procedures) is reversible with a redox potential at ∼405 mV vs Ag/AgCl (Figure 1c), which agrees with the previous report.9 To study the electron-transfer reaction of single Fc-Lip molecules, we bridge a Fc-Lip molecule between the STM tip and substrate electrodes and then sweep the potential while recording the conductance of the molecule (Figure 1d). At low potentials, the redox molecule is in the reduced state, and its conductance is 0.011 ± 0.002 G0, where G0 (= 7.748 × 10−5 S) is the conductance quantum. However, at high potentials, FcLip becomes oxidized, which is observed as a sudden conductance increase by 5× to 0.053 ± 0.007 G0. This finding is consistent with literature6 and further validated by performing a large number of measurements and statistical analysis as discussed below. The measured discrete low and high conductance levels of a single Fc-Lip molecule provide “fingerprinting” for us to determine if the molecule is in the reduced (low conductance) or oxidized (high-conductance)

state, and switching between the two levels measures the individual electron-transfer reaction events in single molecules. Electron-transfer theories assume a coupling between the electronic states and the structure of a molecule and surrounding solvent. Mechanically stretching the molecule is expected to distort the molecular structure and thus the energy levels of the molecule, which could affect electron-transfer reactions. To verify or falsify the idea that there is a link between stretching and changes to redox states, we stretch the molecule by progressively separating the tip and substrate electrodes and monitor the redox state of Fc-Lip from the discrete conductance switching (See Materials and Methods section for details). We perform the measurements by holding the potential at different values: below, at, or above the redox potential of unstretched Fc-Lip (marked on Figure 2a). At potentials well below the redox potential (e.g. −88 mV), Fc-Lip is in its reduced state with a probability of 97% according to the Nernst equation. The individual conductance vs distance traces reveal plateaus corresponding to the formation of single Fc-Lip molecules bridged between the two electrodes, followed by abrupt conductance drops as the bonds at the moleculeelectrode contact break (Figure 2b).10 The conductance plateaus are located near 0.01 G0, which correspond to the low conductance level of the reduced Fc-Lip state. The lack of conductance switching to the high conductance level indicates 14700

DOI: 10.1021/jacs.7b08239 J. Am. Chem. Soc. 2017, 139, 14699−14706

Article

Journal of the American Chemical Society

Figure 2. Mechanical stretching and conductance measurement of Fc-Lip at different potentials. (a) Cyclic voltammogram of Fc-Lip, where the colored dots mark the potentials at which the mechanical stretching and conductance measurements in (b−j) are carried out. Note that the potential is referred to overpotential, E-Eox/red. (b−d) E − Eox/red = −88 mV. The individual conductance traces (b) show plateaus, and the conductance histogram (c) shows a single peak at the low conductance level, indicating that the molecule remains in the reduced state under stretching (d). (e−g) E − Eox/red = +78 mV. The individual conductance traces (e) show plateaus, and the conductance histogram (f) shows a single peak at the high conductance level, indicating that the molecule remains in the oxidized state under stretching (g). (h−j) E - Eox/red = −7 mV, close to the redox potential. The individual conductance traces (h) show plateaus at both the low and high conductance levels and switching from the low to high conductance levels, and the conductance histogram (i) shows two peaks, indicating coexistence of reduced and oxidized species (j).

that the molecule remains in the reduced state during stretching. The corresponding conductance histogram displays a single pronounced peak near 0.01 G0 (Figure 2c), further confirming that mechanical stretching at low potentials does not induce the electron-transfer reaction (Figure 2d). At potentials well above the redox potential, Fc-Lip is expected to be in the oxidized state. For example, at +78 mV, the molecule has a probability of 95% to be in the oxidized state. The individual conductance traces at +78 mV display plateaus around 0.05 G0 (Figure 2e), and the conductance histogram also reveals a single well-defined peak near 0.05 G0

(Figure 2f), showing that the molecule is in the oxidized state during stretching. Like the case where the potential is held well below the redox potential, mechanical stretching at high potentials does not induce electron-transfer reactions (Figure 2g). When the potential is held close to the redox potential, for example, −7 mV, both reduced and oxidized states coexist; this is shown by the presence of plateaus at both the low and high conductance levels in the conductance traces (Figure 2h). At this potential, the conductance often switches abruptly from low to high conductance levels during the stretching of the 14701

DOI: 10.1021/jacs.7b08239 J. Am. Chem. Soc. 2017, 139, 14699−14706

Article

Journal of the American Chemical Society

Figure 3. Mechanical stretching-induced oxidation of single Fc-Lip (E − Eox/red = +30 mV). (a) 2D histogram of conductance plateaus vs stretching distance (counts are normalized for each distance bin), where “R” and “O” mark the reduced and oxidized states. (b−e) 1D conductance histograms for stretching distances between 0 and 0.1 nm (b), 0.1−0.2 nm (c), 0.2−0.3 nm (d), and 0.3−0.4 nm (e), where the red and blue curves are Gaussian fits of the oxidized and reduced conductance peaks, respectively. (f) Logarithmic ratio of the conductance histogram peak area for the oxidized state to that for the reduced state, showing increasing probability of oxidation with stretching distance.

Fc-Lip, we do not observe stretching-induced conductance switching at any potential (Supporting Figure 5b,c), and the conductance histograms display only a single peak (Supporting Figure 5d,e). This control experiment shows that the conductance change in Fc-Lip originates from the ferrocene redox center. We obtain a quantitative relationship between mechanical stretching and redox states by constructing a 2D histogram of conductance vs stretching distance by counting all the traces that show conductance switching at E − Eox/red = +30 mV (Figure 3a). The 2D histogram reveals two conductance regions, corresponding to the low conductance reduced state (“R”) at short stretching distances, and high conductance oxidized state (“O”) at long stretching distances, respectively. 1D conductance histograms for different stretching distances reveal the ratio of the two states and the progressive increase of the oxidized population with stretching (Figure 3b−e). The peak areas are proportional to the numbers (thus probabilities) of Fc-Lip in the reduced and oxidized states13−15(further evidence in Supporting Discussion 1 and Supporting Figure 6). The ratio of the high to low conductance peak areas (including conductance traces with and without switching events (Supporting Figure 7)) is plotted vs stretching distance in Figure 3f, showing that the probability of the molecule in the oxidized state increases with stretching distance. Data measured at +30 mV are shown in Figure 3 as the example to clearly illustrate the stretching effect on the redox reaction. By studying the probabilities of the molecule in the reduced and oxidized states at different potentials and stretching distances, we determine the relationship between mechanical stretching and redox potential (Supporting Figure 8). Figure 4 plots the logarithmic ratios of the probability of the molecule in the oxidized state to that in the reduced state at unstretched

molecule (curves 2 and 3), suggesting that the molecule switches from the reduced to the oxidized state under stretching. The conductance histogram reveals two peaks at 0.013 ± 0.001 G0 and 0.050 ± 0.0011 G0, respectively, corresponding to the reduced and oxidized states, respectively (Figure 2i). We note that previous theoretical studies have revealed decreases in molecular conductance in response to stretching.11,12 In these studies, stretching leads to large decreases in the electronic coupling along the charge transport pathway, via large structural changes, for example, unfolding of the molecule or breakdown of π−π stacking and hydrogen bonding. In our STM experiment, we compare the conductance changes between a fully unfolded molecule and the same structure when additional stretching force is added. Furthermore, stretching leads to a conductance increase in our system, rather than the decrease due to decreased electronic coupling reported in the above references. We also note that our experiments performed at potentials well above and well below the redox potential of Fc-Lip only show a single conductance level with no switching feature (Figure 2b,e), indicating that stretching does not lead to apparent conductance change associated with the electronic coupling. The stretching induced switch-up of conductance only happens to molecules measured at potential close to the redox potential (Figure 2h). The low and high conductance levels are in agreement with the conductance of the reduced and oxidized states (Figure 1d), confirming that the conductance switching is associated with a change in redox state rather than molecular conformation. To further examine the roles of electrode-molecule contact in Fc-Lip, we perform the same measurement on Lip-ctr (control molecule), where the phenyl ring replaces the ferrocene redox center in Fc-Lip (Figure 1a and Supporting Figure 5a). Unlike 14702

DOI: 10.1021/jacs.7b08239 J. Am. Chem. Soc. 2017, 139, 14699−14706

Article

Journal of the American Chemical Society

reduced and oxidized states. Mechanical stretching21,22 shifts the redox potential, given by (Supporting Discussion 2) ΔG0 =

F2 ⎛ 1 1 ⎞ − ⎜ ⎟ − Fq0 2 ⎝ kred kox ⎠

(1)

where F is the force, kred and kox are the effective spring constants of the Cp-Fe-Cp bond in the reduced and oxidized states, respectively, and q0 is the Cp-Fe-Cp bond length increase associated with the oxidation of Fc-Lip. Using F = 1.5 nN for the breakdown force of the molecular junction,10,16,17 q0 = 0.006 nm,19,20 and spring constants determined from the vibrational frequencies by IR spectroscopy23 (νred = 402 cm−1 and νox = 406 cm−1 for the tilting modes) and density functional theory (DFT) calculations (Supporting Table 1 and Supporting Discussion 2), eq 1 predicts −51 mV shift in the redox potential, which is in reasonable agreement with the observation. The model above includes only one coordinate. To consider all possible degrees of freedom, we performed DFT calculations (Figure 5b−e) using the B97-D3 functional24 and an implicit solvation model (See Materials and Methods section for calculation details and see Supporting Methods and Supporting Table 2 for benchmarking study). Under stretching, the molecule undergoes various conformational changes (Figure 5c and Supporting Table 3). Initially, the added strain is accommodated by a slight curvature in the linker side chains. As the stretching increases, the Cp rings begin to twist with respect to each other until the dihedral angle reaches 180°; beyond this point the strain can only be accommodated by stretching bond lengths and angles. The energy changes of the reduced and oxidized states in the twisting and stretching regimes can be fitted by quadratic functions of the stretching distance (Figure 5d, cf. Supporting Figure 9 and Supporting Table 4). The calculations reveal that stretching take places at multiple bonds, but predominantly at Fe−C1 (C9), which is stretched by 0.01 nm at 1.5 nN (Supporting Figure 10). The corresponding adiabatic ionization potential shifts negatively by ∼65 mV (Figure 5e and Supporting Discussion 3). Assuming the thermal and entropic contributions are not affected by stretching, this corresponds to a decrease in the redox potential of ∼65 mV. The value predicted by the analytical model calculation is in good agreement with this more accurate result. It is important to note that unlike the DFT calculations, the simple model calculationd does not consider solvent effects. The good agreements between the simple model and the DFT calculations and also between the calculations and the experimental data indicate that outer sphere/solvent reorganization is not a dominant effect in the present system. This is likely because the STM break junction method measures the last stage of the molecular stretching process. As shown in Figure 5c, the exposure of the iron center to the solvent is not significantly changed (Cp-Fe bond stretched only by 0.01 nm). As such, we believe that change in the outer-sphere reorganization energy with stretching is likely to be relatively small.

Figure 4. Shift of redox potential with mechanical stretching. Logarithmic ratio of the conductance histogram peak area for the oxidized state to that of the reduced state of unstretched (black dots) and stretched (red dots) Fc-Lip vs potential, where the dashed black line is the Nernst equation prediction (with n = 1) for unstretched FcLip. The solid black line is a linear fit to the data for the unstretched molecule, and the solid red line is a linear fit to that for the stretched molecule. In both fits, the slope is 2.303nF/RT with n = 1. The blue arrow marks the shift in the redox potential caused by stretching the molecule over ∼0.3 nm.

(black) and stretched (red) stages. The former refers to data collected during the initial stage of the stretching process with a stretching distance between 0 and 0.1 nm, and the latter corresponds to a fully stretched molecule (0.3−0.4 nm), beyond which the junction breaks down. The breakdown takes place due to the cleavage of an Au−Au bond, which is the weakest bond in the molecular junction with an average breakdown force of ∼1.5 nN.10,16,17 For unstretched redox molecules, the relative probabilities of the molecule in the oxidized and reduced states are expected to follow the Nernst eq (Supporting Discussion 1). Indeed, the data for the unstretched molecule can be fitted by the Nernst equation with a slope of 2.3nF/RT (where n, the number of electron transfer per molecule, is 1, F is the Faraday constant, and R is the gas constant). The data for the stretched molecule are also consistent with the Nernst equation (red line, Figure 4), but the redox potential is shifted negatively by 34 ± 6 mV. The negative shift in the redox potential indicates that the molecule favors the oxidized state under mechanical stretching, which is consistent with the observations in Figures 2 and 3. The data of the stretched state show some deviation from the Nernst equation. Whether this is due to experimental uncertainty or real effect requires further study. It has been found that the coordination bond strengths and lengths for many metal complex are different for the reduced and oxidized states.18 Related to the present work, X-ray crystallography19,20 has shown that the cyclopentadienyl(Cp)Fe-Cp distance in ferrocene increases by 0.006 nm upon oxidation. This is because the electron that is lost upon oxidation is removed from an Fe-Cp bonding orbital, thus weakening the Fe-Cp interaction. We thus expect that mechanically stretching of Fc-Lip will drive the electrontransfer reaction toward oxidation and shift the redox potential negatively. The redox potential shift can be calculated with a single reaction coordinate model (Cp-Fe-Cp bond) in the spirit of the Marcus theory,2 which assumes parabolic energy profiles for the reduced and oxidized states (Figure 5a). In the absence of mechanical stretching, electron transfer occurs when thermal fluctuations bring the molecule to the favorable geometry, corresponding to the interception of the energy profiles of the



CONCLUSIONS Mechanical forces can be an effector of chemical changes22,25 as demonstrated by ball milling/grinding,26,27 ultrasound bath,28,29 and atomic force microscopy,30−33 but measuring a relationship between forces and redistribution of redox equilibria within intact single molecules has been an elusive task. Our study 14703

DOI: 10.1021/jacs.7b08239 J. Am. Chem. Soc. 2017, 139, 14699−14706

Article

Journal of the American Chemical Society

Figure 5. Theoretical model and computational results. (a) Schematic free energy surfaces of the reduced (blue) and oxidized (red) species vs bond length (reaction coordinate) with (dashed curve on the right) and without (solid curve on the left) mechanical stretching. (b) Molecular structure of Fc-Lip showing atomic labeling. (c) DFT predictions of the structural distortions of reduced and oxidized molecules during mechanical stretching. For clarity only the central portion of the molecule is shown. (d) DFT predictions of the potential energy surfaces of the reduced (blue) and oxidized (red) Fc-Lip vs stretching distance (between atoms S1 and S2). Energies are given relative to the global minimum (ring substituents at ∼72°) for each redox state. (e)DFT predictions of changes in the adiabatic ionization potential (aIP) of Fc-Lip vs stretching distance. The aIP is relative to that of unstretched molecule. In (d) and (e), a stretching distance of 0.2793 nm corresponds to an applied force (1.5 nN)that would break the junction (thick dashed line). electrode, respectively. To determine errors in the potential associated with the quasi-reference electrodes, cyclic voltammetry was performed before and after each experiment (Supporting Figure 11), and the measured variation in the redox peak potential of Fc-Lip was taken as errors in the gate voltages. STM break junction experiments were performed with the following two approaches. In the first approach, a small voltage bias (10 mV) was applied between the STM tip and substrate electrode, and the potential was held initially at −0.1 V. The STM tip was brought into contact with the Fc-Lip molecules on the gold substrate, allowing binding of the gold tip to the linker group, and then retracted from the substrate, during which a conductance was monitored. Once a plateau in the conductance was detected, signaling the formation of a single molecule bridged between the tip and substrate, the STM tip was held in position, and the potential was swept at 1 V/s to record the conductance of the molecule vs potential (Figure 1d). In the second approach, both the tip−substrate bias and potential were fixed (see Supporting Table 5 for the applied bias and gate voltages). The STM tip was repeatedly brought into contact and retracted from the substrate repeated, during which the individual conductance traces (conductance vs the STM tip−substrate distance) were collected (Figure 2b,e,h), and a conductance histogram (Figure 2c,f,i) was constructed from the individual conductance traces. The measurement

demonstrates that mechanical stretching induces redox reactions in single metal complex molecules as the stretching changes the bond length and drives the reaction toward the final state that favors the bond length change. This observation supports the thesis that electronic-nuclear coupling is responsible for redox reactions in electron-transfer theories. The observation that a mechanical stretching can control redox states of a single molecule also provides a way to mechanically switch the electronic properties of a single molecule, which supports the vision of molecular machines with electromechanical functions.34



MATERIALS AND METHODS

Measurement of Stretching-Induced Shift in Redox Potential of Single Fc-Lip Molecules. A scanning-tunneling microscope (STM), consisting of a controller (Nanoscope IIIA, Digital Instruments Inc.) and a STM scanner (Molecular Imaging), was used for the break junction measurements. The STM tip was freshly prepared by cutting a gold wire (0.25 mm diameter, 99.95%, Alfa Aesar) and coating with Apiezon wax to reduce the leakage current. The gate voltage was controlled by the bipotentiostat (Agilent). A silver wire and a platinum coil were used as quasi-reference electrode and counter 14704

DOI: 10.1021/jacs.7b08239 J. Am. Chem. Soc. 2017, 139, 14699−14706

Journal of the American Chemical Society



was repeated at differential potentials. Conductance vs distance histograms (2D) were also obtained from the conductance vs distance traces (Supporting Figure 7a), from which 1D conductance histograms at different stretching distances were obtained (Supporting Figure 7c− f). Computational Methods. The changes in adiabatic ionization potential (aIP) of the Fc-Lip with stretching and the associated geometric distortions were explored using DFT calculations performed with Gaussian 09.35 After an extensive benchmarking study (see Supporting Methods and Supporting Table 2), the B97-D324 functional was chosen to describe the system, along with the def2svp36 basis sets for C, H, O, and S atoms and the more extensive def2tzvp36 basis set for Fe. Solvent effects (water) were included using the SMD37 implicit solvation model. Relaxed potential energy surface (PES) scans were performed for both the reduced and oxidized species, with the distance between atoms S1 and S2 being scanned (Supporting Figure 9), and all other geometric coordinates being allowed to fully relax. Only conformations where the side chains were fully extended were considered, as the redox potential changes are expected to be primarily associated with strain on the ferrocene moiety and comparatively insensitive to the conformations of the alkyl chains. This was confirmed for a smaller model compound. To ensure the lowest energy electronic configuration was obtained for the oxidized species, the optimized orbitals from the minimum energy structure were used as an initial guess for each new geometry when scanning the PES. Scans were performed iteratively for both stretching and relaxation of strain until a smooth curve was obtained. A harmonic curve was fitted to the PES data to obtain the force constant (second derivatives of the curves). These could be used to determine the degree of stretching the system could undergo before sufficient force had been applied to cleave an Au−Au bond (1.5 nN).10,16,17 Due to mechanical stretching the molecule, the data points do not correspond to stationary points on the PES. Additionally, due to the fixed ends of the linker, the system cannot undergo free rotation and translation. As a result, standard statistical mechanics formulas (which are derived for an ideal gas) cannot be applied to calculate the thermal and entropic effects on this system, and hence the free energy and redox potential changes as the molecule is stretched cannot be determined. Instead, only the response of the aIP (in the presence of a solvent field) to stretching is reported. We note that this would be equivalent to the change in redox potential if the thermal and entropic effects are unaffected by increasing strain.



Article

AUTHOR INFORMATION

Corresponding Authors

*[email protected] *[email protected] *[email protected] ORCID

Yueqi Li: 0000-0001-7003-1978 Naomi L. Haworth: 0000-0002-3299-3137 Michelle L. Coote: 0000-0003-0828-7053 Nongjian Tao: 0000-0002-5206-153X Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors would like to thank Drs. Wolfgang Schmickler and Andrew Gilbert for stimulating discussions. Financial support from the Office of Naval Research (N00014-11-1-0729) and from the Australian Research Council (CE140100012 and DE160100732) and generous allocations of supercomputing time on the National Facility of the Australian National Computational Infrastructure are gratefully acknowledged.



REFERENCES

(1) Kuznetsov, A. M.; Ulstrup, J. Electron transfer in chemistry and biology. An introduction to the theory; Wiley: Chichester, 1999. (2) Marcus, R. A. Annu. Rev. Phys. Chem. 1964, 15, 155. (3) Skourtis, S. S.; Waldeck, D. H.; Beratan, D. N. Annu. Rev. Phys. Chem. 2010, 61, 461. (4) Schmickler, W.; Rampi, M. A.; Tran, E.; Whitesides, G. M. Faraday Discuss. 2004, 125, 171. (5) Nitzan, A.; Ratner, M. A. Science 2003, 300, 1384. (6) Xiao, X.; Brune, D.; He, J.; Lindsay, S.; Gorman, C. B.; Tao, N. Chem. Phys. 2006, 326, 138. (7) Xu, B.; Tao, N. J. Science 2003, 301, 1221. (8) Li, Y.; Xiang, L.; Palma, J. L.; Asai, Y.; Tao, N. Nat. Commun. 2016, 7, 11294. (9) Zanello, P. Inorganic Electrochemistry, Theory, Practice and Application; The Royal Society of Chemistry: Cambridge, UK, 2003. (10) Huang; Chen, F.; Bennett, P. A.; Tao. J. Am. Chem. Soc. 2007, 129, 13225. (11) Lin, J.; Beratan, D. N. J. Phys. Chem. A 2004, 108, 5655. (12) Franco, I.; George, C. B.; Solomon, G. C.; Schatz, G. C.; Ratner, M. A. J. Am. Chem. Soc. 2011, 133, 2242. (13) Artés, J. M.; López-Martínez, M.; Díez-Pérez, I.; Sanz, F.; Gorostiza, P. Small 2014, 10, 2537. (14) Arielly, R.; Vadai, M.; Kardash, D.; Noy, G.; Selzer, Y. J. Am. Chem. Soc. 2014, 136, 2674. (15) Xiang, L.; Palma, J. L.; Li, Y.; Mujica, V.; Ratner, M. A.; Tao, N. Nat. Commun. 2017, 8, 14471. (16) Rubio, G.; Agraït, N.; Vieira, S. Phys. Rev. Lett. 1996, 76, 2302. (17) Xu, B.; Xiao, X.; Tao, N. J. J. Am. Chem. Soc. 2003, 125, 16164. (18) Hao, X.; Zhu, N.; Gschneidtner, T.; Jonsson, E. Ö .; Zhang, J.; Moth-Poulsen, K.; Wang, H.; Thygesen, K. S.; Jacobsen, K. W.; Ulstrup, J.; Chi, Q. Nat. Commun. 2013, 4, 2121. (19) Dunitz, J. D.; Orgel, L. E.; Rich, A. Acta Crystallogr. 1956, 9, 373. (20) Martinez, R.; Tiripicchio, A. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1990, 46, 202. (21) Bell, G. I. Science 1978, 200, 618. (22) Stauch, T.; Dreuw, A. Chem. Rev. 2016, 116, 14137. (23) Duggan, D. M.; Hendrickson, D. N. Inorg. Chem. 1975, 14, 955. (24) Grimme, S.; Ehrlich, S.; Goerigk, L. J. Comput. Chem. 2011, 32, 1456. (25) James, S. L.; Adams, C. J.; Bolm, C.; Braga, D.; Collier, P.; Friscic, T.; Grepioni, F.; Harris, K. D. M.; Hyett, G.; Jones, W.; Krebs,

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.7b08239. Supporting experimental methods and DFT calculation methods. Supporting discussions on (1) the r elationship between conductance peak area and surface coverage of redox species; (2) derivation of eq 1, spring constant calculation, and further estimation; and (3) theoretical description of the stretching process. Supporting figures of NMR spectra, control experiment, plateau length statistics, analysis of both switched and unswitched curves at +30 mV, peak ratio vs stretching distance at all potentials, calculated potential energy surface vs distance, calculated significant structural parameters vs stretching distance, and calibration of quasi-reference electrode. Supporting tables of spring constants and shifts in Gibbs free energy, benchmark study of theoretical calculation, calculated molecular geometries, calculated potential energy surface, applied overpotential in each experiments (PDF) 14705

DOI: 10.1021/jacs.7b08239 J. Am. Chem. Soc. 2017, 139, 14699−14706

Article

Journal of the American Chemical Society A.; Mack, J.; Maini, L.; Orpen, A. G.; Parkin, I. P.; Shearouse, W. C.; Steed, J. W.; Waddell, D. C. Chem. Soc. Rev. 2012, 41, 413. (26) Boldyrev, V. In Treatise on Materials Science and Technology; Herbert, H., Ed.; Academic Press: New York, 1983; Vol. 19, Part B, p 185. (27) Balaz, P.; Achimovicova, M.; Balaz, M.; Billik, P.; CherkezovaZheleva, Z.; Criado, J. M.; Delogu, F.; Dutkova, E.; Gaffet, E.; Gotor, F. J.; Kumar, R.; Mitov, I.; Rojac, T.; Senna, M.; Streletskii, A.; Wieczorek-Ciurowa, K. Chem. Soc. Rev. 2013, 42, 7571. (28) Basedow, A. M.; Ebert, K. H. In Physical Chemistry; Springer: Berlin, Heidelberg, 1977; p 83. (29) Hickenboth, C. R.; Moore, J. S.; White, S. R.; Sottos, N. R.; Baudry, J.; Wilson, S. R. Nature 2007, 446, 423. (30) Rief, M.; Gautel, M.; Oesterhelt, F.; Fernandez, J. M.; Gaub, H. E. Science 1997, 276, 1109. (31) Wiita, A. P.; Perez-Jimenez, R.; Walther, K. A.; Grater, F.; Berne, B. J.; Holmgren, A.; Sanchez-Ruiz, J. M.; Fernandez, J. M. Nature 2007, 450, 124. (32) Baldus, I. B.; Gräter, F. Biophys. J. 2012, 102, 622. (33) Dopieralski, P.; Ribas-Arino, J.; Anjukandi, P.; Krupicka, M.; Kiss, J.; Marx, D. Nat. Chem. 2013, 5, 685. (34) Browne, W. R.; Feringa, B. L. Nat. Nanotechnol. 2006, 1, 25. (35) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, O.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09; Gaussian, Inc.: Wallingford CT, 2009. (36) Weigend, F.; Ahlrichs, R. Phys. Chem. Chem. Phys. 2005, 7, 3297. (37) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. J. Phys. Chem. B 2009, 113, 6378.

14706

DOI: 10.1021/jacs.7b08239 J. Am. Chem. Soc. 2017, 139, 14699−14706